<<

arXiv:math/0111287v1 [math.AT] 27 Nov 2001 htwe optn oooyclmt o oooia pcsoen true. one A but spaces Strange, in (given issue. topological theorem cofibrancy for this curious colimits about a worry prove computing we when this around that get To cofibrant. ,btfrnww iljs iea nutv ento.A pnhyperco open An definition. intuitive an give just will we now for but 4, hproes,rte hnjust than rather ‘hypercovers’, ro htaod ea’ hoe opeey n sqieelementar ro quite this is going and of completely, instead theorem Segal’s But detecte avoids are unity. that equivalences of proof weak partitions that have fact always deduc the spheres to of possible use is It making covers. result, open arbitrary for equivalence weak o n ipiilobject simplicial any For otntl hs rtraapyol hnteobjects the 1 when Th. [H, only (cf. apply colimit criteria homotopy these its with fortunately agrees realization geometric its eaedfiieyntasmn htteoe sets open the that assuming not definitely are we u rtga nti ae st eeaieti eutt h olwn t following the to result this generalize to 1.1. is Theorem paper this in goal first Our sawa equivalence. weak a is | Here ehv hsnntt rwtedgnrce o yorpia r typographical for degeneracies the if draw that to proved not [S1] chosen have we oytheory. topy hsdt n a ul the build may one data this C ˇ ( h anga fti ae sgnrlzn hoe . ota tapplies it that so 1.1 Theorem generalizing is paper this of goal main The h eodse st elwt h ieec between difference the with deal to is step second The hr r w tp nteagmn.Frt epoethat prove we First, argument. the in steps two are There 1991 Date Let e od n phrases. and words Key U h eodato a upre ya S ototrlResea Postdoctoral NSF an by supported was author second The ) U ∗ X a → | oebr2,2001. 25, November : X ahmtc ujc Classification. Subject 0 Abstract. cee vrra n ope fields. complex and real over schemes ocntuttplgclraiainfntr o the for realization topological construct to h aua a hocolim map natural the ··· sasmlca space simplicial a is eatplgclsae n let and space, topological a be a n X = sa is U o vr pncover open every For a 0 homotopy eso htif that show We ···∩ · ∩· AILDGE N AILC ISAKSEN C. DANIEL AND DUGGER DANIEL X YECVR NTOPOLOGY IN HYPERCOVERS ` a atto fuiysbriaeto subordinate unity of partition a has U yecvr oooyclmt emti elzto,mot realization, geometric colimit, homotopy hypercover, W a U 0 a ∗ n o o ehcomplex Cech namdlctgr,teeaegnrlciei o when for criteria general are there category, model a in ˇ n h aemp r bandb mtigindices— omitting by obtained are maps face the and , qiaec,where equivalence, U U ∗ ehcvr.Teeaedfie ndti nSection in detail in defined are These covers. Cech ˇ ∗ 1. uhthat such U → ` ∗ Introduction sahproe fatplgclspace topological a of hypercover a is X U U a 53,1F0 14F42. 14F20, 55U35, sawa qiaec.Ti ati used is fact This equivalence. weak a is 0 a of 1 1 U o o o X = h aua map natural the , C ˇ { ( U ` U |−| a ) } U ∗ U ea pncvrof open an be eoe emti realization. geometric denotes a hc stesmlca space simplicial the is which , 0 a A a | 1 1 n hi nescin are intersections their and C a hmtp hoyof theory -homotopy W ˇ 2 ( n U · · · c Fellowship. rch ) r l obat and cofibrant, all are ∗ | hocolim n hocolim and | hsfo Segal’s from this e C yshrs and spheres, by d ˇ U ( U X aos Segal easons. hntemap the then y. ) t egv a give we ute ∗ then C pni A) ppendix → | vrhsto has ever ˇ ..];un- 9.6.4]); heorem. ( X U vchomo- ivic e fa of ver ) From . ∗ C X ˇ → ( U sa is ) X to ∗ . 2 DANIELDUGGERANDDANIELC.ISAKSEN

(1) Each Un is a disjoint of open of X, (2) The spaces appearing in U0 are an open cover of X, (3) The spaces in U1 cover the double intersections of those in level 0, (4) The spaces in U2 cover the triple intersections of those in level 1, and so on. Of course making sense of (4)—especially the ‘and so on’ part—requires a certain amount of bookkeeping, which is why we are postponing the formal definition. But the essence is that hypercovers are like Cechˇ complexes except that instead of taking the double intersections at level 1 we may refine them further, and we may continue this refining process at each level. Our second main result is then

Theorem 1.2. If U∗ is an open hypercover of a space X, then the natural map hocolim U∗ → X is a weak equivalence. This result could almost be considered folklore since everyone immediately agrees it’s true, but a proof seems to be missing from the literature. One might consider tackling it by appealing to the Whitehead theorem, proving an isomorphism on fundamental groupoids and with local coefficients. This is the approach taken in [F, Prop. 8.1] in the related context of ´etale hypercovers, but this is messy and obscures in computation the underlying geometric explanation of the theorem. In the case of topological spaces, the isomorphism on fundamental groupoids was the subject of the paper [RT] (although they only dealt with Cechˇ complexes, not hypercovers). The approach we take here, on the other hand, is very elementary. The idea is to reduce to the case of Cechˇ covers in a clever way. Our interest in these results arose from attempts to understand topological re- alization functors in the A1- of schemes [MV]. Given an algebraic variety X defined over C, there is an associated X(C) obtained by giving X the analytic . Of course this should extend to a map of ‘homotopy theories’ from the Morel-Voevodsky category Spc(C) to the category of topological spaces. In [MV] this extension is only provided at the level of homotopy categories, but we are interested in extending it to the model category level. The key fact needed to make this work is precisely Theorem 1.2. This is worked out in detail in Section 5, following the basic program of [I] (also outlined in [D2, Rem. 8.2]). We also prove that taking analytic spaces for schemes defined over R induces a Quillen map from Spc(R) to Z2-equivariant topological spaces. Finally, we give in this paper several interesting corollaries to Theorem 1.2. On the whole these seem too disparate to recount in the introduction, but as an example let us mention two of them. We refer the reader to Sections 3 and 4 for more results like these. Corollary 1.3. Let E → B be any map which is locally split (for example, a ), and form the associated Cechˇ complex Cˇ(E)∗ given by ˇ n+1 C(E)n := EB = E ×B E ×B ···×B E.

Then the natural map hocolim Cˇ(E)∗ → B is a weak equivalence.

Corollary 1.4. Let U be an open cover of a space X with the property that every U finite intersection Ua0···an is covered by other elements of . Form the diagram consisting of all the Ua’s and all the inclusions between them. Then the homotopy colimit of this diagram is weakly equivalent to X. HYPERCOVERS IN TOPOLOGY 3

The first corollary is an immediate consequence of Proposition 4.10, and the second is restated and proved as Proposition 4.6(c). Using open covers to give homotopy decompositions for spaces, or to detect weak equivalences, is of course a classical topic. In addition to [S1] it is worthwhile to mention [Mc1], [Mc2], and [Dk]. Hypercovers were invented by Verdier in [SGA4, Expose V, Sec. 7], where they were used as a way of computing in arbitrary Grothendieck . We would like to express our thanks to Bill Dwyer, Phil Hirschhorn, Michael Mandell, and Jeff Smith for several useful conversations about these results. 1.5. Notation, terminology, and other annoyances. We assume that the reader is familiar with homotopy colimits, and in a few places also with the theory of model categories. The original reference for the latter is [Q], but we generally follow [H] in notation and terminology ([Ho] is also a good reference). Regarding homotopy colimits, [H] uses ‘hocolim D’ to denote the result of applying a certain explicit formula to any diagram D. This has the disadvantage that the resulting object has the correct homotopy type only when the diagram consists entirely of cofibrant objects. We instead adopt the position that ‘hocolim D’ should always denote the correct homotopy-invariant construction: it is obtained by first applying cofibrant-replacement to the objects in the diagram, and only then using the usual explicit formulas. In model-theoretic terms, homotopy colimit is the left derived of the ordinary colimit functor, when the category of diagrams is given the projective model structure (see below). Having made the previous , we now get to say that for topological spaces it isn’t really necessary. This is definitely a non-standard fact, but we’ve banished it to Appendix A so it won’t distract the reader from the general theme of the paper. On the other hand, it is a useful result and we’d like to call the reader’s attention to it: when taking homotopy colimits for diagrams of topological spaces one doesn’t first have to make all the spaces involved cofibrant. The usual formulas are already homotopy-invariant. We review one last piece of machinery, used often in the body of the paper. Given a small category I, recall that there is a model structure on the category of diagrams sSetI such that a map is a weak equivalence (resp., fibration) if it is so in every spot of the diagram [H, Sec. 13.8]. We call this the projective model structure on sSetI , and the cofibrant diagrams have the property that the homotopy colimit and ordinary colimit are weakly equivalent. Finally, some notation: Throughout this paper our open covers U = {Ua} are always indexed by a A. In particular, we are allowing the possibility that ′ Ua = Ua′ for different values a 6= a . For every finite set σ = {a0,... ,an} in A, we’ll write Uσ or Ua0···an for Ua0 ∩···∩ Uan . Also, once and for all we fix our model n n+1 for ∆ as the of R consisting of (n + 1)-tuples t = (t0,... ,tn) such that n T 0 ≤ ti ≤ 1 for all i and Σi=0ti = 1. The symbol op denotes the category of all topological spaces—we don’t assume any hypotheses like compactly-generated.

2. Cechˇ complexes The purpose of this section is to prove the following: Theorem 2.1. For any open cover U of a topological space X, the natural map π : |Cˇ(U)∗| → X is a weak equivalence. 4 DANIELDUGGERANDDANIELC.ISAKSEN

We start by recalling the following result and its corollary: Proposition 2.2 (Gray). Let f : X → Y be a map of spaces and let U and V form an open cover of Y . Suppose that the induced maps f −1U → U, f −1V → V, and f −1(U ∩ V ) → U ∩ V are all weak equivalences. Then X → Y is also a weak equivalence. This is proven (in more generality) in [Gr, 16.24], using an elegant small-simplices argument. With enough technology it can also be done by a Whitehead-type theo- rem: it’s easy to see that X → Y is an isomorphism on π0, a souped-up van Kampen theorem yields the isomorphism on π1, and for homology with local coefficients one uses the Mayer-Vietoris . Gray’s argument is much nicer, though.

Corollary 2.3 (May). Let f : X → Y be a map of spaces and let U = {Ua} be an −1 open cover of Y . Suppose that f Uσ → Uσ is a weak equivalence for every finite set σ of indices. Then X → Y is also a weak equivalence. May deduces the generalization by a quick application of Zorn’s Lemma [M2, −1 Cor. 1.4]: look at the set of all opens W such that f (W ∩ Uσ) → W ∩ Uσ is a weak equivalence for all σ, including σ = ∅. This set has a maximal element, and Gray’s result shows it must be X. In an earlier paper McCord proved a more general version of this result [Mc1, Th. 6], but the proof is quite a bit more complicated. Proof of Theorem 2.1. Given any V in X, the space π−1(V ) is homeomor- ′ ′ phic to the space |Cˇ(U )∗|, where U is the open cover {Ua ∩ V } of the space V . This definitely uses the fact that V is open. −1 ′ We want to consider the maps π (Uσ) → Uσ, but in this case the cover U of Uσ actually contains the whole space Uσ as one of its elements. From the following ′ lemma we know that under this condition |Cˇ(U )∗| → Uσ is a weak equivalence; so by Corollary 2.3 the map |Cˇ(U)∗| → X is a weak equivalence as well.

Lemma 2.4. Let U be an open cover of X such that Ub = X for some index b. Then the natural map |Cˇ(U)∗| → X is a weak equivalence (in fact, a homotopy equivalence). 0 Proof. There is a section χ: X → |Cˇ(U)∗| obtained from the map Ub ⊗ ∆ → |Cˇ(U)∗| and the identification Ub = X. We only need to show that χπ is homotopic to the identity. Let Cˇ(U)∗ ×I be the simplicial space obtained by crossing all the levels of Cˇ(U)∗ with the unit . Then |Cˇ(U)∗ × I| is the quotient

n Ua0···an × ∆ × I / ∼ "a0···a # a n where the relations are the usual ones, not affecting the I factor at all. Define a map |Cˇ(U)∗ × I| → |Cˇ(U)∗| in the following way. Take an element (x, t0,... ,tn,s) n where x belongs to Ua0···an and (t0,... ,tn) belongsto ∆ , and send it to the element n+1 (x, 1 − s,st0,... ,stn) in the factor Uba0...an ⊗ ∆ . This definition respects the various identifications. Now, there is also an obvious map f : |Cˇ(U)∗ × I| → |Cˇ(U)∗|× I induced by sending (x,t,s) to ((x, t),s). We claim that this is a , thereby giving us a homotopy |Cˇ(U)∗|× I → |Cˇ(U)∗| between χπ and the identity. The reason f HYPERCOVERS IN TOPOLOGY 5 is a homeomorphism is just because geometric realization and crossing with I are both left adjoints, and the right adjoints are easily seen to commute. It is important that I and ∆n are locally compact Hausdorff so that the relevant mapping spaces with compact-open topologies have the correct adjointness properties. 2.5. Connection with Segal’s results. To close this section we make the con- nection between our Theorem 2.1 and the result proven in [S1]. Segal doesn’t explicitly deal with Cechˇ complexes, but the objects he deals with turn out to be homeomorphic to them. This connection will be needed later on. Let A be the indexing set for a cover U. We have already introduced the Cechˇ complex Cˇ(U)∗, but if A is given an ordering we may also consider the ordered o Cechˇ complex Cˇ (U)∗ which is often easier to work with. This is the simplicial ˇo U space given by C ( )n = a0···an Ua0···an , where the coproduct ranges over all ordered multi-indices in A. That is, we only consider multi-indices for which a0 ≤ ` o a1 ≤···≤ an. Note that there is an inclusion of simplicial spaces Cˇ (U)∗ → Cˇ(U)∗. o Proposition 2.6. The map Cˇ (U)∗ → Cˇ(U)∗ induces a homotopy equivalence o |Cˇ (U)∗| → |Cˇ(U)∗|.

Proof. For any (not necessarily ordered) multi-index a0 ··· an, there is a canonical reordering aσ0 ··· aσn such that aσ0 ≤···≤ aσn. If ai = aj for some i < j, then o always choose σi < σj. This allows us to define an inverse map |Cˇ(U)∗| → |Cˇ (U)∗|. n If (x, t) is an element of Ua0···an ⊗ ∆ , then send (x, t) to the element (x,σt) of n Uaσ0···aσn ⊗ ∆ , where σt is defined by (σt)i = tσi. One composition is the equal to the identity. It remains to construct a homotopy H : |Cˇ(U)∗|× I → |Cˇ(U)∗| between the other composition and the identity. As in the proof of Lemma 2.4, we use the space |Cˇ(U)∗ × I| rather than |Cˇ(U)∗|× I. n ˇ U We define H as follows: An element (x, t) of Ua0···an ⊗ ∆ is equivalent in |C( )∗| 2n+1 to the element (x, t0,... ,tn, 0,..., 0) of Ua0···anaσ0···aσn ⊗ ∆ . Also, (x,σt) is ˇ U equivalent in |C( )∗| to the element (x, 0,..., 0,tσ0,... ,tσn) of Ua0···anaσ0···aσn ⊗ ∆2n+1. Define H((x, t),s) to be the element

(x,st0,... ,stn, (1 − s)tσ0,..., (1 − s)tσn) 2n+1 of Ua0···anaσ0···aσn ⊗ ∆ .

Proposition 2.7. Let U be an open cover of a space X indexed by a set A. Con- sider the realization of the simplicial space

[n] 7→ Uσn , σ0⊆···⊆σ a n where the coproduct is indexed by chains of nonempty, finite subsets of A. This o realization is homeomorphic to the realization |Cˇ (U)∗| of the ordered Cechˇ complex and is homotopy equivalent to |Cˇ(U)∗|. The realization in the above proposition is the object considered in [S1]. The ordered Cechˇ complex is another construction of the same space, which for us seems somewhat easier to work with. One disadvantage, of course, is that it is not natural: a total ordering on A must be chosen to begin with. Proof. The second claim follows from the first claim and Proposition 2.6. For the first claim, it is convenient to use a slightly unusual construction of o |Cˇ (U)∗|. When forming the geometric realization, instead of forming Cartesian 6 DANIELDUGGERANDDANIELC.ISAKSEN products with ∆k we instead form products with sd ∆k; since they are homeomor- phic it doesn’t matter which one we use. Given this, the key observation is that we can coordinatize sd ∆k in the following way: assuming that the vertices of ∆k are labelled by the numbers 0,... ,k in the usual way, a point on sd ∆k is represented uniquely by a chain of proper inclusions σ0 ⊂ ··· ⊂ σj of subsets of {0,...,k} together with an element t of ∆j . Essentially, the chain of subsets determines in which sub- the point lies, and then t gives local coordinates inside that sub-simplex. Using this coordinate , we can write down maps in both directions be- tween the two realizations

n k Uσn × ∆ / ∼ and Ua0···ak × sd ∆ / ∼ . " σ0⊆···⊆σ # " a0≤···≤a # a n a k For instance, let’s give the map from left to right. Using degeneracy relations, a point p in the left space can be represented by a chain of proper inclusions n σ0 ⊂··· ⊂ σn, a point x of Uσn , and an element t of ∆ . Let a0,a1,... ,ak be the ordered list of elements of σn. The chain σ∗ together with t defines a point s in sd ∆k, and so we map p to the pair (x, s). It is easy to see that this map is well-defined and continuous, and just as easy to write down its inverse.

In the case that {Ua} admits a partition of unity {ψa} it is fairly easy to see o that the map π : |Cˇ (U)∗| → X admits a section: First, a point x of X has a neighborhood which intersects the of ψa only for finitely many indices o a = a0,... ,an. The section χ sends x to the point of |Cˇ (U)∗| represented by (x, t) n in Ua0···an ⊗ ∆ where ti = ψai (x). One has to check that χ is continuous (use the local-finiteness of the partition of unity), and that χπ ≃ id via a straight-line homotopy. See Proposition 4.1 of [S1].

3. Passing to homotopy colimits The results of the previous section all concerned geometric realizations. In this section we translate these into results about various homotopy colimits. In gen- eral, there is a ‘Reedy cofibrancy’ condition on simplicial spaces which guarantees that geometric realization and homotopy colimit agree. Unfortunately our Cechˇ complexes are not Reedy cofibrant, due to the fact that the open sets appearing in them are not necessarily cofibrant spaces. However, Theorem A.8 shows that in the category of topological spaces this cofibrancy issue is unimportant: homotopy colimits can be computed naively, without first making things cofibrant. This fact saves the day.

Theorem 3.1. If U is an open cover of a space X, then the natural map hocolim Cˇ(U)∗ → X is a weak equivalence. Proof. By Theorem A.8, we can compute the homotopy colimit in the Strom model category. In this model structure the Cechˇ complex is Reedy cofibrant (it has free degeneracies in the sense of Definition A.4), and so the realization already has the correct homotopy type. Theorem 2.1 now gives the result. Here are several alternative formulations: HYPERCOVERS IN TOPOLOGY 7

Proposition 3.2. Let A be an indexing set for the cover U, and let PA denote the partially ordered set consisting of all nonempty finite subsets of A. Let Γ denote Pop T the functor A → op which sends σ to Uσ. Then the natural map hocolim Γ → X is a weak equivalence. Proof. To construct hocolim Γ we can take the realization of the simplicial replace- ment for Γ (by Theorem A.8 we don’t need to first make the spaces cofibrant). That is, we take the realization of the simplicial space

[n] 7→ Uσn , σ0⊆···⊆σ a n where the coproduct is indexed by chains of nonempty, finite subsets of A. Now Proposition 2.7 tells us that this realization is homotopy equivalent to |Cˇ(U)∗|, so Theorem 2.1 finishes the proof.

Corollary 3.3. Let PU denote the subcategory of Top whose objects are the open sets Ua belonging to U together with their finite intersections; the morphisms are the inclusions of open subsets of X. Let Γ denote the inclusion functor PU → Top. Then the natural map hocolim Γ → X is a weak equivalence. Pop P Proof. Consider the obvious functor F : A → U sending σ to Uσ. We will show that it is homotopy cofinal, so pick an object V in PU and look at the undercategory (V ↓ F ). It suffices to show that any map K → N(V ↓ F ) can be extended over the cone on K, as K ranges over all finite simplicial sets. Every n-simplex s in

K maps to a chain of open sets V → Uσ0 → Uσ1 → ··· → Uσn in (V ↓ F ). Since K has only finitely-many non-degenerate simplices, only finitely-many of the Uσ will ever appear. Define µ to be the union of all the σi arising from the map K → N(V ↓ F ). To extend the map over CK, we send the cone on s to the (n+1)- simplex corresponding to the chain V → Uµ → Uσ0 → Uσ1 →···→ Uσn .

The following corollary was shown to us by Bill Dwyer. Let (Top ↓ X)U denote the full subcategory of (Top ↓ X) consisting of all maps Z → X that factor through T T the space E = a Ua. Let Γ: ( op ↓ X)U → op be the canonical functor sending Z → X to Z. We would like to claim that the homotopy colimit of the diagram Γ ` is weakly equivalent to X, but (Top ↓ X)U is not a small category. So we choose an infinite cardinal κ larger than the size of E and restrict to the spaces Z that have at most κ elements. As the proof of the corollary indicates, the weak homotopy type of hocolim Γ is independent of the choice of κ, as long as κ is sufficiently large.

Corollary 3.4. For the functor Γ: (Top ↓ X)U → Top defined above, the natural map hocolim Γ → X is a weak equivalence. ˇ n Proof. The nth level of the Cech complex is EX := E ×X E ×X ···×X E (n factors). op Let’s write C = (Top ↓ X)U, for brevity. So we have the functor F : ∆ → C given n op C T ˇ U by [n] 7→ EX . The composition ∆ → → op is just C( )∗. Because of Theorem 3.1, it will be enough to show that F is homotopy cofinal. For this we pick an object z : Z → X in C and show that (z ↓ F ) is contractible. This undercategory is isomorphic to the category of simplices of K, where K is the n n sending [n] to HomC(z, EX ). But observe that HomC(z, EX ) is equal n n to T where T = HomC(z, EX ). So K is the simplicial set [n] 7→ T , which is contractible because T is nonempty (using the fact that z : Z → X factors through 8 DANIELDUGGERANDDANIELC.ISAKSEN

E). Thus (z ↓ F ) is isomorphic to the category of simplices of a contractible simplicial set, and therefore has a contractible nerve.

Corollary 3.5 (Small simplices theorem). Let Sing UX denote the simplicial set n whose n-simplices are the maps ∆ → X that factor through some Ua. Then Sing UX → Sing X is a weak equivalence.

Proof. Let PA be the category defined in Proposition 3.2, where A is the indexing Pop S set for the cover. Consider the diagram Γ: A → s et defined by Γ(σ)= Sing (Uσ). By general nonsense |hocolim Γ| ≃ hocolim |Γ|. Also, there is a commutative dia- gram

hocolimPop |SingU | / A σ |Sing X|

  hocolimPop U / A σ X in which the vertical maps are weak equivalences because the natural map |SingY | → Y is a weak equivalence for every space Y . We know from Proposition 3.2 that the bottom horizontal map is a weak equivalence, so the top horizontal map is also a weak equivalence. We conclude that the map hocolim Γ → Sing X is a weak equivalence of simplicial sets. Therefore, we shall compare hocolim Γ and Sing UX. Pop For the , assume that A is finite. Notice that A is a Reedy category [Ho, Def. 5.2.1], where we think of all the maps as being directed upward. Since there are no non-identity downward maps, the fibrations are objectwise in the Reedy model Pop structure on sSet A (see [Ho, Th. 5.2.5]). So in this case the Reedy and projective model structures (cf. Section 1.5) are the same. In particular, a Reedy-cofibrant diagram is also projective-cofibrant, which guarantees that the homotopy colimit and the ordinary colimit are weakly equivalent. The functor Γ may be checked to be Reedy cofibrant: at the spot indexed by σ = {a0,... ,an}, the latching object is the subobject of SingUσ consisting of all simplices which are contained in some other Ub. The fact that it is actually a subobject says that the latching map is a cofibration. So we know that hocolim Γ and colim Γ are weakly equivalent. It is easy to check that colim Γ =∼ Sing UX. We have shown that if U is a finite cover, then Sing UX is weakly equivalent to Sing X. Now let A be arbitrarily large. For any finite subcollection U′, let ∪U′ denote the ′ ′ ′ union of the open sets in U . Then we know the map Sing U′ (∪U ) → Sing (∪U ) is a weak equivalence. But Sing UX → SingX is the filtered colimit of these maps, where the indexing category is the poset of all finite subcollections U′. This uses that each space ∆n is compact. Our result now follows from the fact that filtered colimits of simplicial sets preserve weak equivalences.

4. Theorems In this section we define hypercovers, and then prove our main result, Theo- rem 1.2. We go on to deduce various corollaries. Before giving a rigorous definition of hypercovers, we need to recall a few pieces of machinery related to simplicial objects. For any category C, let sC denote the category of simplicial objects in C. Likewise, let s≤nC denote the category of HYPERCOVERS IN TOPOLOGY 9 truncated simplicial objects of dimension n. There is the obvious forgetful functor skn : sC → s≤nC, and if C has all finite limits then skn has a right adjoint called coskn; these are the skeleton and coskeleton functors. If U∗ belongs to sC then we’ll often abbreviate coskn(skn U)∗ as just coskn U∗. Finally, the nth matching object MnU is defined to be the nth object of coskn−1 U∗. There is a canonical map of simplicial spaces U∗ → coskn−1 U∗, and in level n it gives Un → MnU. In X levels less than n, this map is the identity. We write coskn for the nth coskeleton functor for s(Top ↓ X). These definitions have somewhat easier interpretations when C is the category of topological spaces. To describe these, note that any simplicial set may be regarded as a simplicial space which is discrete in every dimension, and if U∗ and W∗ are simplicial spaces then the set of maps from U∗ to W∗ has a natural topology coming from the compact-open topology on spaces. Using these observations, one checks that n (i) Un =∼ Map(∆ ,U∗), k (ii) [coskn U]k =∼ Map(skn∆ ,U∗), and n (iii) MnU =∼ Map(∂∆ ,U∗). The first property is immediate from the Yoneda lemma. The second property follows from the first and the adjunction between skn and coskn. The third property is a special case of the second. Finally, say that a map of spaces Z → X is an open covering map if it is isomorphic to a map of the form a Ua → X where {Ua} is an open cover of X. Definition 4.1. A hypercover `of a space X is an augmented simplicial space X U∗ → X such that the maps Un → Mn U are open covering maps for all n ≥ 0. Here X T Mn U denotes the nth matching object of U∗ computed in the category s( op ↓ X) of simplicial spaces over X. X Note that M0 U =∼ X, so the condition for n = 0 says that U0 → X is an X ∼ open covering map. Also M1 U = U0 ×X U0, so when n = 1 we are requiring U1 → U0 ×X U0 to be an open covering map. The reader should be aware that X when n > 1 the objects MnU and Mn U turn out to be isomorphic, so one can forget about the extra complication of the overcategory. Using properties (i)–(iii) above, it can be checked that if U∗ → X is a hypercover and K → L is an inclusion of finite simplicial sets, then the map Map(L,U∗) → X Map(K,U∗) is also an open covering map. From this, it follows that coskn U∗ → X X is a hypercover whenever U∗ → X is a hypercover. Also, each map Uk → [coskn U]k is an open covering map. We leave it to the reader to check that in a hypercover each Un must be a disjoint union of open subsets of X, and that Cechˇ complexes are the hypercovers for X which the maps Un → Mn U are all isomorphisms. Generalizing this, a hypercover X U∗ → X is called bounded if there exists an N such that the maps Un → Mn U are isomorphisms for all n>N. The smallest such N for which this happens is called the dimension of the hypercover. Said intuitively, the bounded hypercovers of dimension N are the hypercovers for which the refinement process stops after the Nth level. A hypercover U∗ → X has dimension at most N if and only if ∼ X U∗ = coskN U∗.

Lemma 4.2. If U∗ → X is a bounded hypercover, then hocolim U∗ → X is a weak equivalence. 10 DANIELDUGGERANDDANIELC.ISAKSEN

A more detailed version of the following proof, given in the context of an arbitrary , appears in [DHI]. Proof. We proceed by induction, starting from the fact that bounded hypercovers of dimension 0 are just Cechˇ covers and therefore are handled by Theorem 3.1. Suppose that U∗ → X is a bounded hypercover of dimension n + 1. Define V∗ to be coskn U∗, so V∗ is a bounded hypercover of dimension at most n. Therefore, we may assume by induction that hocolim V∗ → X is a weak equivalence. The canonical map U∗ → V∗ gives an open covering map Un+1 → Vn+1, by the very definition of what it means for U∗ to be a hypercover (since Vn+1 = Mn+1U). In fact, one can check that Uk → Vk is an open covering map for all k. Consider the following bisimplicial object, augmented horizontally by V∗:

o o o V∗ U∗ o U∗ ×V∗ U∗ o ···

The kth row is the (augmented) Cechˇ complex for the open covering map Uk → Vk. Note that for 0 ≤ k ≤ n the kth row is the constant simplicial object with value Uk because Uk → Vk is the identity. Call this bisimplicial object (without the horizontal augmentation) W∗∗. Let D∗ denote the diagonal of W∗∗. Standard homotopy theory tells us that hocolim D∗ may be computed (up to weak equivalence) by first taking the homotopy colimits of the rows of W∗∗, and then taking the homotopy colimits of the resulting simplicial object. But the homotopy colimit of the kth row is just Vk by Theorem 3.1. Since V∗ is a bounded hypercover of dimension at most n, we have assumed that hocolim V∗ is weakly equivalent to X. So hocolim D∗ → X is a weak equivalence. We claim that U∗ is a retract, over X, of D∗. Note first that one has, in complete generality, a map U∗ → D∗; in dimension k it is the unique horizontal degeneracy W0k → Wkk. To produce a map D∗ → U∗ it is enough to give skn+1 D∗ → skn+1 U∗, because U∗ = coskn+1 U∗. Notice that skn D∗ = skn U∗. Choosing any face map [0] → [n+1] gives a map Wn+1,n+1 → W0,n+1, which is just Dn+1 → Un+1. This induces a corresponding map skn+1 D∗ → skn+1U∗ as desired. It is straightforward to check that U∗ → D∗ → U∗ is the identity (because U∗ = coskn+1 U∗ one only has to check it on (n + 1)-skeleta), and all the maps commute with the augmentations down to X. We have already shown that hocolim D∗ → X is a weak equivalence. Since hocolim U∗ → X is a retract of hocolim D∗ → X, it must also be a weak equivalence.

Theorem 4.3. If U∗ → X is a hypercover then the maps hocolim U∗ → |U∗| → X are all weak equivalences.

Proof. The fact that hocolim U∗ → |U∗| is a weak equivalence follows just as in Theorem 3.1 for the case of Cechˇ complexes: we may compute the homotopy colimit in the Strom model category, where the simplicial object U∗ is Reedy cofibrant since it has free degeneracies (Definition A.4). To show that |U∗| → X is a weak equivalence, note first that we have an iso- morphism πk|U∗| → πk|coskk+1 U∗|. (This is true for any map of simplicial spaces X∗ → Y∗ which is an isomorphism on (k +1)-skeleta—an easy proof is to apply the singular functor everywhere to get into bisimplicial sets, then use the diagonal in HYPERCOVERS IN TOPOLOGY 11 place of realization.) But coskk+1 U∗ is a bounded hypercover, so Lemma 4.2 tells =∼ us that πk|coskk+1 U∗| −→ πkX. 4.4. Complete covers. In this section we don’t quite consider hypercovers, but rather a related concept which captures the same phenomena. This second approach was suggested to us by Jeff Smith.

Definition 4.5. An open cover U = {Ua} of a space X is called complete if for all finite sets σ of indices, the intersection Uσ is covered by elements of U. It is called a Cechˇ cover if every Uσ is again an element of the cover. Complete covers appear in [DT, Satz 2.2], where they were used in the context of identifying quasi-fibrations. The paper [Mc1] then used them to detect weak equivalences. We blur the distinction between a cover and the full subcategory that it spans inside the category of open sets of X. Given a cover U, we can construct an associated simplicial space in the following way: For any n ≥ 0, let Pn denote the category of nonempty subsets of {0,... ,n}, where the maps are the inclusions. Note that the assignment [n] 7→ Pn defines a cosimplicial category in the obvious way. (Application of the nerve functor everywhere gives the cosimplicial space [n] → sd ∆n.) Define Ω∗ to be the simplicial space [n] 7→ F ({0,...,n}), : op U F Pan → op U where the coproduct runs over all functors Pn → . The faces and degeneracies are induced by those in P in the expected way. To give a point in Ω3, for example, is to give the following data: (1) A sequence of opens U0,...,U3 in U, (2) 6 open subsets U01,U02,...,U23 in U such that Uij ⊆ Ui ∩ Uj; (3) 4 open subsets U012,...,U123 in U such that Uijk ⊆ Uij ∩ Ujk ∩ Uik; (4) An open subset U0123 in U which is contained in all the Uijk; (5) A point on U0123. It is usually helpful to think of these open sets as indexed by the faces of a 3-simplex. In forming the Cechˇ complex of a cover U we are throwing in all the finite intersections Uσ into the higher levels of the simplicial object, and these are typically objects which are not in U itself. The simplicial object Ω∗ is in some sense the closest thing we can get to a Cechˇ complex while requiring all the open sets to belong to U. Proposition 4.6.

(a) If the cover U is complete then Ω∗ is a hypercover of X. (b) Regarding U as a category, let Γ: U → Top be the obvious inclusion. Then hocolim Γ ≃ |Ω∗|. (c) If the cover U is complete then the natural map hocolim Γ → X is a weak equivalence.

Proof. For part (a), consider the full subcategory P¯n of Pn consisting of all objects except for {0, 1,...,n}. Then the matching space MnΩ is equal to F¯(σ) . ¯ ¯op U ¯ F : Pan → hσ\∈Pn i 12 DANIELDUGGERANDDANIELC.ISAKSEN

For example, a point in M3Ω is determined by the data in (1)–(3) above, together with a point in U012 ∩ U013 ∩ U023 ∩ U123. ¯ ¯op U Since the cover is complete, for each functor F : Pn → and each element op ¯ ¯ ¯ x of ∩σ∈Pn F (σ), there exists an extension F of F to Pn such that x belongs to F ({0,...,n}). This shows that Ωn → MnΩ is an open covering map, which finishes part (a). Part (b) is almost trivial, given the right machinery. To form hocolim Γ we can work in the Strom model structure on Top (see Appendix A), where we first take the simplicial replacement

[n] 7→ U0

U0→···→U a n and then form the realization. Here the coproduct is indexed over all functors n n ∆ → U, where ∆ denotes the category of n composable maps. Note that Ω∗ was formed in almost the same way as the simplicial replacement of Γ, except we indexed op U the coproduct by functors Pn → . Each Pn is essentially just a subdivision of n ∆ , so it’s not surprising that |Ω∗| is another model of the homotopy colimit. In somewhat more detail: Let sd′ denote the ‘opposite’ of the usual subdivision functor on sSet, in which the orientations of all the simplices have been changed so that they point away from the barycentres, rather than towards them. (We op ′ need this because we are using Pn rather than Pn.) The functor sd has a right adjoint Ex′. There is a natural ‘first vertex map’ sd′ K → K, inducing K → Ex′ K. Given our diagram Γ: U → Top, the realization of the simplicial replacement is isomorphic to the coend Γ ⊗U B, where B : U → sSet sends Ua to the classifying space B(Ua ↓ U). Likewise, one checks that the realization of Ω∗ is isomorphic to ′ ′ the coend Γ ⊗U Ex B, where Ex B is the obvious composite functor. The natural U map B → Ex′ B is an objectwise weak equivalence. The object B of sSet is cofibrant (see [H, Cor. 15.8.8]), where this diagram category has the projective model structure described in Section 1.5. The exact same arguments show that Ex′ B is also cofibrant in this structure. So we have an objectwise weak equivalence between two cofibrant diagrams. The diagram Γ: U → Top is objectwise cofibrant (since we are working with the Strom model structure on Top), and so by [H, ′ Cor. 19.3.5] it follows that Γ ⊗U B → Γ ⊗U Ex B is a weak equivalence. Finally, part (c) is an immediate consequence of (a), (b), and Theorem 4.3. The following corollary was originally proven by McCord [Mc1, Th. 6], but is an easy consequence of our hypercovering theorem. It generalizes May’s result from Corollary 2.3, which handled the case of Cechˇ covers. For the proof we will need the following observations: (1) If U → X is an open covering map and f : Y → X is any map, there is an induced open covering map Y ×X U → Y . (2) If U∗ → X is −1 a hypercover and f : Y → X is a map of spaces, one gets a hypercover f U∗ → Y whose space in level n is Y ×X Un. Corollary 4.7. Let f : X → Y be a map of spaces. Suppose there is a complete −1 cover U = {Ua} of Y such that each f (Ua) → Ua is a weak equivalence. Then f itself is a weak equivalence. U Y Proof. From form the associated hypercover Ω∗ as described in the paragraph X −1 Y preceding Proposition 4.6. Pulling this back to X gives a hypercover Ω∗ := f Ω∗ , as described above (note that this is not the hypercover associated to the covering −1 X Y {f Ua}). Now f induces a map Ω∗ → Ω∗ compatible with the augmentations. HYPERCOVERS IN TOPOLOGY 13

This map of simplicial spaces is a levelwise weak equivalence, by assumption. Upon taking homotopy colimits we get X ∼ / Y hocolim Ω∗ hocolim Ω∗

∼ ∼   X / Y, and so we conclude that X → Y is also a weak equivalence. 4.8. Generalized hypercovers for topological spaces. Up until now we have only considered open covers, but now we turn to a broader notion. We’ll say that a map p: E → B of spaces is a generalized cover if it is locally split: that is, every element of B has a neighborhood U such that p−1(U) → U admits a section. Observe that covering spaces, and in fact fibre bundles in general, are generalized covers. The point for us is that generalized covers and open covers generate the same Grothendieck topology on topological spaces.

Definition 4.9. An augmented simplicial space U∗ → X is a generalized hyper- X cover of X if the maps Un → Mn U are generalized covers.

Proposition 4.10. If U∗ is a generalized hypercover of X then hocolim U∗ → X is a weak equivalence. Proof. Results from [DHI], in the context of an arbitrary Grothendieck topology, show that this is a consequence of Theorem 4.3. The essential point is that gener- alized hypercovers can all be refined by open hypercovers. To deduce this from the results of [DHI] we do the following: Pick a regular T cardinal λ larger than the size of all the sets in U∗. Let opλ denote the category of topological spaces of size less than λ, and make it into a Grothendieck site via T the usual notion of open cover. Form the universal model category U( opλ) (see the paper [D2]) and localize it with respect to the set S consisting of all maps hocolim V∗ → X, where V∗ → X is an open hypercover. Theorem 4.3 implies that T T that there is a ‘realization map’ U( opλ)/S → op. The results of [DHI] say that T in U( opλ)/S one actually knows that hocolim U∗ → X is a weak equivalence for all generalized hypercovers U∗, and then applying our realization functor tells us this must hold in Top as well. Corollary 1.3 is an immediate consequence of the above proposition. Example 4.11. Let G be a topological and consider the usual covering space ξ : EG → BG. Form the Cechˇ complex Cˇ(ξ)∗, which is a generalized hypercover of BG. Using only the fact that EG has a free G-action, one can see that the nth level n of Cˇ(ξ)∗ is homeomorphic to G × EG, and the face and degeneracy maps are the familiar ones of the two-sided bar construction B(∗, G, EG). Now using that EG is contractible, we find that Cˇ(ξ)∗ is levelwise weakly equivalent to the simplicial space o o ∗ o G oo G × G ··· .

The above proposition tells us that |Cˇ(ξ)∗| ≃ BG, and so in this way we recover the usual bar construction for BG. 14 DANIELDUGGERANDDANIELC.ISAKSEN

5. Topological realization functors for A1-homotopy theory Let k be a field. Morel and Voevodsky [MV] produced a model category Spc(k) which captures the ‘motivic homotopy theory’ of smooth schemes over k. Here Spc(k) stands for ‘spaces over k’. It is the category of simplicial presheaves on the Nisnevich site of smooth schemes over Spec k. When k comes with an k ֒→ C, then any k-scheme X gives rise to a topological space X(C) consisting of its C-valued points with the analytic topology. A natural expectation is to use this functor to relate Spc(k) to the usual model category Top of topological spaces. Morel and Voevodsky showed how to extend this functor on the level of homotopy categories (by somewhat awkward methods), but they didn’t produce functors at the model category level. In this section we use Proposition 4.10 to produce such functors, with the small provision that we have to replace Spc(k) with a Quillen-equivalent variant. We also address the situation when k ֒→ R, in which case one can construct topological realization functors into Z2-equivariant spaces. As in [D2], a Quillen pair L: M ⇄ N: R will be called a Quillen map M → N.

5.1. The Complex case. Let T denote either the Zariski, Nisnevich, or ´etale Grothendieck topology on the category Sm/k of smooth k-schemes. In the termi- ′ nology of [D2], let Spc (k)T denote the universal model category built from Sm/k subject to the following relations: ∼ (1) X ∐ Y −→ (X ∪ Y ) (here ∐ denotes the coproduct in our model category, whereas ∪ denotes disjoint union of schemes); ∼ (2) hocolim U∗ −→ X for any T-hypercover U∗ of a smooth scheme X (called ‘basal hypercovers’ in [DHI]); ∼ (3) X × A1 −→ X. (Relation (1) is morally a special case of (2), but must be included separately for technical reasons—see [DHI]). ′ The model categories Spc(k)T and Spc (k)T have the same underlying category and the same class of weak equivalences, but differ in their notions of cofibration and fibration. They are injective and projective versions of the same homotopy theory. ′ ′ Theorem 5.2. There are Quillen maps Spc (k)et → Top and Spc (k)Nis → Top sending a smooth k-scheme X to X(C). ′ Proof. By general nonsense from [D2], to give a Quillen map Spc (k)T → Top we just need to give a functor Sm/k → Top which respects the above relations. The functor we’re interested in is X 7→ X(C), and this clearly preserves relations (1) and (3). In the case of the ´etale topology, the fact that it preserves relation (2) is just Proposition 4.10; the point is that if p: E → B is an ´etale cover, then p(C): E(C) → B(C) satisfies the hypotheses of the inverse function theorem and hence is locally split. Since the ´etale topology is finer than the Nisnevich topology, there is an obvious ′ ′ map Spc (k)Nis → Spc (k)et (in essence, there are more relations of type (2) for the ′ ´etale topology). So one also gets a topological realization map Spc (k)Nis → Top by composition. It is possible to show that the functor X 7→ X(C) takes elementary distinguished squares [MV] to homotopy pushout squares of topological spaces. Together with HYPERCOVERS IN TOPOLOGY 15 results of [B], this can be used to give an alternative proof of the above theorem for the Nisnevich topology. The Real case. If we have a Real field k ֒→ R, then the space X(C) comes .5.3 equipped with an action of the group Gal(C/R)= Z2. So we might hope to compare ′ Spc (k) to a model category of Z2-equivariant spaces. Recall that if G is a finite group then there are two notions of weak equivalence for G-spaces, called the non-equivariant and G-equivariant equivalences. An equivariant map X → Y is a non-equivariant equivalence if it is a weak equivalence after forgetting the equivariant structure, and it is a G-equivariant equivalence if XH → Y H is a non-equivariant weak equivalence for every subgroup H ⊆ G. There are associated G-equivariant and non-equivariant model structures on the category of G-spaces, which we will denote Top(G) and Top(G)non. If p: E → B is an equivariant map which is also a covering space (non- equivariantly), the map hocolim Cˇ(E)∗ → B is a non-equivariant equivalence but not necessarily a G-equivariant equivalence. For instance, if p is G → ∗ then the map hocolim Cˇ(E)∗ → B is equal to EG →∗. So when we have a subfield k ֒→ R the arguments given above show that the functor X 7→ X(C) induces a Quillen map ′ ′ Spc (k)et → Top(Z2)non, but not a Quillen map Spc (k)et → Top(Z2). However, when we use the Nisnevich topology something special happens. Lemma 5.4. If E → B is a Nisnevich cover of k-schemes, then E(C)Z2 → B(C)Z2 is locally split. For a counterexample to this in the case of ´etale covers, try Spec C → Spec R. Proof. First note that X(C)Z2 is homeomorphic to X(R) for any scheme X over k. The map p(R): E(R) → B(R) is surjective by the defining property of Nisnevich covers; every R-point in B lifts to E. By definition of ´etale covers, p(R) satisfies the hypothesis of the inverse function theorem. Since p(R) is surjective, it is locally split.

′ Theorem 5.5. There is a Quillen map Spc (k)Nis → Top(Z2) sending a smooth k-scheme X to X(C). Proof. The argument exactly parallels the non-equivariant case in Theorem 5.2, so the only nontrivial part is to show that if U∗ → X is a Nisnevich hypercover then the map hocolim U∗(C) → X(C) is a Z2-equivariant weak equivalence of Z2- spaces. The fact that it is a non-equivariant equivalence has already been discussed in Theorem 5.2, because U∗ → X is in particular an ´etale hypercover. So we must consider what happens when we take Z2-fixed points. It is a fact that for any diagram D of G-spaces (G any finite group) and any subgroup H of G, one has (hocolim D)H ≃ (hocolim DH ) (see Remark 5.6 below). Z2 Z2 So we just need to convince ourselves that hocolim{U∗(C) } → X(C) is a non- Z2 equivariant weak equivalence. But by the above lemma one sees that U∗(C) is a generalized hypercover of X(C)Z2 , and so the result is an instance of Proposi- tion 4.10.

Remark 5.6. In the above proof we needed the fact that (hocolim D)H is weakly equivalent to hocolim(DH ). This is well-known in equivariant topology, but it’s hard to find an actual reference. We give a brief sketch, for which we are grateful to Michael Mandell. 16 DANIELDUGGERANDDANIELC.ISAKSEN

First of all, it clearly suffices to consider the case where all the Di are cofibrant. This means in particular that they are Hausdorff. We form hocolim D by first writing down the simplicial replacement of the diagram, and then taking geometric realization. Taking H-fixed points obviously commutes with the simplicial replace- ment functor, so it suffices to worry about the geometric realization part. But one can check that if X∗ is a simplicial space in which all Xn are Hausdorff, then H H |X∗| is homeomorphic to |X∗ |. To do this, use the skeletal filtration on |X∗| and the fact that |Skn X∗| is obtained from |Skn−1 X∗| by pushing out along a closed inclusion (this is one of the places where the Hausdorff condition is needed). Check that taking fixed-points commutes with filtered colimits, and for Hausdorff spaces it also commutes with pushouts along closed inclusions.

Appendix A. Homotopy colimits for diagrams of non-cofibrant spaces Let Top denote the category of all topological spaces, with its usual model cate- gory structure. Given a diagram D : I → Top, the usual instructions for computing the homotopy colimit of D are (1) to apply a cofibrant-replacement functor to every object in the diagram, and (2) to then use an explicit formula like that of Bousfield-Kan [BK, Sec. XII.2]. This is the situation in an arbitrary model category. In this section we show that for the special case of Top, the first step of cofibrant- replacement is actually not needed. What we show is that no matter what formula one uses for computing homotopy colimits—whether it is the Bousfield-Kan for- mula or your favorite alternative—that formula always gives a homotopy invariant construction in Top, even without the cofibrant-replacement step. This fact seems not to be well known, although it could be argued that the seeds lie there in the collective subconscious of algebraic topologists. In any case, for our purposes here we need to bring it into the light of day. The most useful way to formulate this result seems to be in model category terms, as a comparison between the usual model structure on Top and the Strom model structure, where everything is cofibrant. See Theorem A.8. We would like to thank Phil Hirschhorn for helpful conversations about the results in this section, in particular for his ideas on removing an annoying T1 sep- aration condition. The final form of Lemmas A.2 and A.3 is something we owe to him.

To begin with, we need the following Lemma A.1. Let A → B and X → Y be weak equivalences. Given a diagram A × Dn o A × Sn−1 / X

   B × Dn o B × Sn−1 / Y, where the maps in the left-hand-square are the obvious ones, the induced map from the pushout of the top row to the pushout of the bottom row is also a weak equiva- lence. Note that if A and B are cofibrant then this is an easy consequence of left- properness for Top, but we claim the result in greater generality. HYPERCOVERS IN TOPOLOGY 17

Proof. Let XA and YB be the pushouts of the top and bottom rows respectively, and write f : XA → YB for the map between them. We will produce a suitable cover of these spaces and use Proposition 2.2. Let UB be the pushout of B × (Dn −{0}) o B × Sn−1 / Y. n Write Dǫ for {x ∈ D : |x| < ǫ} (where 0 <ǫ< 1), and let VB = B × Dǫ. The spaces UB and VB clearly form an open cover of YB , and notice that UB deformation- retracts down to Y . The intersection UB ∩ VB is equal to B × (Dǫ −{0}). The same definitions give us a cover {UA, VA} of XA, and it is easy to check −1 −1 −1 that f (UB) = UA and f (VB ) = VA. So the map f (VB ) → VB is the map A × Dǫ → B × Dǫ, which is a weak equivalence. Similar reasoning shows −1 that f (UB ∩ VB) → UB ∩ VB is a weak equivalence. Finally, one argues that −1 f (UB) → UB is a weak equivalence because it deformation-retracts down to X → Y . Proposition 2.2 now shows that XA → YB is a weak equivalence.

We’ll say that an inclusion Y ֒→ Z is relatively T1 if given any open set U in Y and any point z of Z\U, there is an open set W of Z such that U ⊆ W and z∈ / W (compare the similar definition from [Ho, p. 50]). It follows that if E is any finite subset of Z\U, one can find an open set W ⊆ Z which contains U and doesn’t intersect E. Note that a space X is T1 precisely if all the inclusions {x} ֒→ X are relatively T1. Lemma A.2. Given a pushout diagram of the form A × Sn / Y  

  A × Dn+1 / Z,

.the inclusion Y ֒→ Z is relatively T1 Proof. Suppose given a point z in Z and an open U in Y . Either z is in Y or else it is represented by a pair (a,t) where t is in the of Dn+1. The argument works the same for the two cases, and so for convenience we’ll assume the latter. n Pull back U to A × S and express it as a union of rectangles Vi × Wi, where Vi n is open in A and Wi is open in S . Each Wi can be fattened into an open subset ′ n+1 ′ n ′ Wi of D with the properties that Wi ∩ S = Wi and Wi does not contain t. ′ n+1 Let M be the union of the Vi × Wi ; it is an open subset of A × D . Let N be the union of the images of M and U in Z. One checks that N ∩ Y = U, and the pullback of N to A × Dn+1 is M. So N is open in Z and N contains U, but N does not contain z.

The following lemma is well-known for closed inclusions of T1-spaces (see also [Ho, Prop. 2.4.2]). The usual proof still works in our case.

Lemma A.3. Suppose that Y1 ֒→ Y2 ֒→ · · · is a sequence of relatively T1 inclusions and that K is a . Then any map f : K → colim Y factors through some Yk.

Proof. Suppose the map does not factor through any Yk. By taking a subsequence of Y if necessary, we can find a sequence of points k1, k2,... in K with the property that f(ki) lies in Yi\Yi−1. 18 DANIELDUGGERANDDANIELC.ISAKSEN

Pick an n and set Vn = Yn. Next, choose an open set Vn+1 in Yn+1 which contains Vn but doesn’t contain f(kn+1). Then pick an open set Vn+2 in Yn+2 which contains Vn+1 but neither f(kn+1) nor f(kn+2). Continuing this process gives an infinite sequence of opens, so their colimit Wn is an open subset of colim Y . As n varies, the open subspaces Wn form a cover of colim Y . But f(K) is a compact subspace of colim Y , and it is not covered by any finite subcover. This is a contradiction. We now need some machinery related to simplicial spaces.

Definition A.4. A simplicial space X∗ is said to be split, or to have free degen- eracies, if there exist subspaces Nk ֒→ Xk such that the canonical map

Nσ → Xk σ a is an isomorphism. Here the variable σ ranges over all surjective maps in ∆ of the form [k] → [n], Nσ denotes a copy of Nn, and the map Nσ → Xk is the one induced ∗ by σ : Xn → Xk (see [AM, Def. 8.1]).

The idea is that the spaces Nk represent the ‘non-degenerate’ part of Xk, sitting inside of Xk as a direct summand. It is an easy exercise to check that if X∗ has free degeneracies and all the Nk are cofibrant spaces, then X∗ is Reedy cofibrant in sTop. If X∗ is any simplicial space, let Skn X∗ be the simplicial space equaling X∗ through dimension n and equaling the degenerate subspaces of X∗ in larger di- mensions. This is slightly different than the n-truncated simplicial space skn X∗. There are maps Sk0 X∗ → Sk1 X∗ → · · · and the colimit is X∗. It follows that |X∗| is equal to colimn |Skn X∗|, using that geometric realization is a left adjoint (and this doesn’t require any assumptions on X, only hinging upon the fact that the spaces ∆n are locally compact Hausdorff). An important point is that when X∗ has free degeneracies the space |Skn X∗| is obtained from |Skn−1 X∗| via the pushout diagram n / (A.1) Nn × ∂∆ |Skn−1 X∗|

  n / Nn × ∆ |Skn X∗|.

Proposition A.5. Let X∗ be a simplicial space with free degeneracies. If K is a compact space then any map K → |X∗| factors through some |Skn X∗|. Proof. This is a direct application of Lemmas A.2 and A.3, using the skeletal fil- tration of |X∗| and the pushout square (A.1). The following corollary is the crucial ingredient for Theorem A.8. It is very similar to things in the literature, notably [M1, Th. 11.13] and [S2, Lem. A.5]. May’s result assumes the spaces are compactly-generated and Hausdorff, and also that the realizations are simply-connected. Segal’s result is more similar to ours, and the proofs follow the same pattern, but he works with homotopy equivalences rather than weak equivalences. HYPERCOVERS IN TOPOLOGY 19

Corollary A.6. If X∗ → Y∗ is a map of simplicial spaces with free degeneracies such that Xn → Yn is a weak equivalence for each n, then |X∗| → |Y∗| is also a weak equivalence.

Proof. For every k and every basepoint ∗ of X0, there is an isomorphism

colim πk(| Skn X∗|, ∗) → πk(|X∗|, ∗) n

(and the same statement holds with X∗ replaced by Y∗). This follows from Proposition A.5, taking K to be a sphere. Therefore, it suffices to show that |Skn X∗| → |Skn Y∗| is a weak equivalence. Using induction, this follows from the pushout square (A.1) and Lemma A.1. Recall that the Strom model category is a model structure for topological spaces, denoted TopS , in which the weak equivalences are homotopy equivalences and the cofibrations (resp., fibrations) are the Hurewicz cofibrations (resp., fibrations). Note that all objects are cofibrant in this structure. Proposition A.7. The Strom model category is left proper and simplicial. Proof. Left properness is automatic when all objects are cofibrant [H, Cor. 11.1.3]. The simplicial action is of course given by A ⊗ K =∼ A × |K|. To establish the simplicial structure we use the reductions outlined in [D1, Sec. 3]. If A → B is a Hurewicz cofibration and K → L is a cofibration of simplicial sets, then |K| → |L| is a closed cofibration and therefore the map A ⊗ L B ⊗ K → B ⊗ L A⊗K a ∼ is a cofibration by [L, Cor. 1]. If A ֌ B is a Hurewicz cofibration and a homotopy equivalence, then certainly A ⊗ K → B ⊗ K is still a homotopy equivalence. And ∼ if K ֌ L is a trivial cofibration of simplicial sets then |K| → |L| is actually a homotopy equivalence, hence A ⊗ K → A ⊗ L is also a homotopy equivalence.

Theorem A.8. Let D : I → Top be a diagram of spaces. Then the homotopy colimits of D as computed in Top and TopS have the same weak homotopy type. Proof. In TopS, since all objects are cofibrant, we can compute hocolim D by first taking the simplicial replacement of D and then applying the realization functor. In Top we first apply a cofibrant-replacement functor to all the objects in the diagram, and only then do we take simplicial replacement and realize. Simplicial replacements always have free degeneracies (see [D2, Proof of Lem. 2.7]), hence Corollary A.6 applies.

Remark A.9. Theorem A.8 also holds if one uses the category of compactly- generated, weak Hausdorff spaces with its usual model structure. The same proofs work, with some extra caution that the various colimits are what they’re supposed to be. 20 DANIELDUGGERANDDANIELC.ISAKSEN

References

[SGA4] M. Artin, A. Grothendieck, and J.L. Verdier, Th´eorie des Topos et Cohomologie Etale des Sch´emas, Springer Lecture Notes in Math. 270, Springer-Verlag, Berlin, 1972. [AM] M. Artin and B. Mazur, Etale´ homotopy, Springer Lecture Notes in Math. 100, Springer- Verlag, Berlin, 1969. [B] B. Blander, Local projective model structures on simplicial presheaves, K-theory, to ap- pear. [BK] A.K. Bousfield and D.M. Kan, Homotopy limits, completions, and localizations, Springer Lecture Notes in Math. 304, Springer-Verlag, New York, 1972. [Dk] T. tom Dieck, Partitions of unity in homotopy theory, Comp. Math. 23, Fasc. 2 (1972), 159–167. [DT] A. Dold and R. Thom, Quasifaserungen und unendliche symmetrische produkte, Ann. Math. 67, no. 2, (1958), 239–281. [D1] D. Dugger, Replacing model categories by simplicial ones, Trans. Amer. Math. Soc. 353 (2001), 5003-5027. [D2] D. Dugger, Universal homotopy theories, Adv. Math., to appear. [DHI] D. Dugger, S. Hollander, and D. C. Isaksen, Hypercovers and simplicial presheaves, preprint. [F] E. M. Friedlander, Simplicial schemes and ´etale homotopy type, Annals of Mathematics Studies 104, Princeton University, 1982. [Gr] B. Gray, Homotopy Theory: An Introduction to , Academic Press, 1975. [H] P. S. Hirschhorn, Localization of Model Categories, preprint, version dated April 12, 2000. (Available at http://www-math.mit.edu/∼psh). [Ho] M. Hovey, Model Categories, Mathematical Surveys and Monographs vol. 63, Amer. Math. Soc., 1999. [I] D. C. Isaksen, Etale realization on the A1-homotopy theory of schemes, preprint. [L] J. Lillig, A union theorem for cofibrations, Arch. Math. XXIV (1973), pp. 410–414. [M1] J.P. May, Geometry of iterated loop spaces, Springer Lecture Notes in Math. 271, Springer-Verlag, Berlin-New York, 1972. [M2] J.P. May, Weak equivalences and quasifibrations, Groups of self-equivalences and related topics (Proc. Conf. Universit´ede Montr´eal, Montreal, Quebec, 1988), Lecture Notes in Math. 1425, Springer-Verlag, Berlin, 1990, pp. 91–101. [Mc1] M.C.McCord, Singular homology groups and homotopy groups of finite topological spaces, Duke Math. J. 33 (1966), 465–474. [Mc2] M.C. McCord, Homotopy type comparison of a space with complexes associated with its open covers, Proc. Amer. Math. Soc. 18 (1967), 705–708. [MV] F. Morel and V. Voevodsky, A1-homotopy theory of schemes, Inst. Hautes Etudes Sci. Publ. Math. 90 (2001), 45–143. [Q] D. Quillen, Homotopical Algebra, Springer Lecture Notes in Math. 43, Springer-Verlag, Berlin, 1969. [RT] A. Razak Salleh and J. Taylor, On the relation between fundamental groupoids of the classifying space of the nerve of an open cover, J. Pure Appl. Algebra 37 (1985), no. 1, 81–93. [S1] G. Segal, Classifying spaces and spectral sequences, Inst. Hautes Etudes Sci. Publ. Math. 34 (1968), 105–112. [S2] G. Segal, Categories and cohomology theories, Topology 13 (1974), 293–312.

Department of Mathematics, Purdue University, West Lafayette, IN 47907

Department of Mathematics, University of Notre Dame, Notre Dame, IN 46556 E-mail address: [email protected] E-mail address: [email protected]